This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On Lie algebras of generalized Jacobi matrices

Alice Fialowski University of Pécs and Eötvös Loránd University
Budapest, Hungary
E-mail: [email protected], [email protected]
   Kenji Iohara Univ Lyon, Université Claude Bernard Lyon 1
CNRS UMR 5208, Institut Camille Jordan,
F-69622 Villeurbanne, France
E-mail: [email protected]
Abstract

In these lecture notes, we consider infinite dimensional Lie algebras of generalized Jacobi matrices 𝔤J(k)\mathfrak{g}J(k) and 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k), which are important in soliton theory, and their orthogonal and symplectic subalgebras. In particular, we construct the homology ring of the Lie algebra 𝔤J(k)\mathfrak{g}J(k) and of the orthogonal and symplectic subalgebras.

keywords:
Infinite dimensional Lie algebras, Lie algebra Homology, Cyclic and Hochschild Homology, smash product, spectral sequence.
\mathclass

Primary 17B65, 16S35; Secondary 16E40.

\abbrevauthors

A. Fialowski and K. Iohara \abbrevtitleGeneralized Jacobi matrices

\maketitlebcp

1 Introduction

In this expository article, we consider a special type of general linear Lie algebras of infinite rank over a field kk of characteristic zero, and compute their homology with trivial coefficients.

Our interest in this topic came from an old short note (2 pages long) of Boris Feigin and Boris Tsygan (1983, [FT]). They stated some results on the homology with trivial coefficients of the Lie algebra of generalized Jacobi matrices over a field of characteristic 0. It was a real effort to piece out the precise statements in that densely encoded note, not to speak about the few line proofs. In the process of understanding the statements and proofs, we got involved in the topic, found different generalizations and figured out correct approaches. We should also mention that their old results generated much interest during these 36 years. In the meantime, two big problems have beed solved, which helped us to understand the statements and work out the right proofs. One is the Loday-Quillen-Tsygan theorem (see (3.1)) and the other is the generalization of the Hochschild-Serre type spectral sequence by Stefan (see (3.4)), without which the statements of the old note could not be justified.

Consider the Lie algebra 𝔤J(k)\mathfrak{g}J(k) of generalized Jacobi matrices, namely, infinite size matrices M=(mi,j)M=(m_{i,j}) indexed over \mathbb{Z} such that mi,j=0m_{i,j}=0 if |ij|>N|i-j|>N for some NN depending on the matrix MM. The original Jacobi operator, also known as Jacobi matrix, is a symmetric linear operator acting on sequences which is given by an infinite tridiagonal matrix (a band matrix that has nonzero elements only on the main diagonal, the first diagonal below this, and the first diagonal above the main diagonal.) It is commonly used to specify systems of orthonormal polynomials over a finite, positive Borel measure. This operator is named after Carl Gustav Jacob Jacobi, who introduced in 1848 tridiagonal matrices and proved the following Theorem: every symmetric matrix over a principal ideal domain is congruent to a tridiagonal matrix (see [K] and e.g. [Sz]). Since then, Jacobi matrices play an important role in different branches of mathematics, like in topology (Bott’s periodicity theorem on homotopy groups), stable homotopy theory, algebraic geometry and CC^{*}-algebras (see [Ka1]). They are also used to show some interesting properties in K-theory. For instance, Karoubi used them to prove the conjecture of Atiyah-Singer about the classifying space, see [Ka2].

The Lie algebra 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k) of finitely supported infinite size matrices indexed over \mathbb{Z} can be naturally viewed as a subalgebra of 𝔤J(k)\mathfrak{g}J(k).

The Lie algebra 𝔤J(k)\mathfrak{g}J(k) has typical infinite dimensional nature. For example, the two matrices

P:=iiEi1,i,Q:=iEi+1,i,P:=\sum_{i\in\mathbb{Z}}iE_{i-1,i},\qquad Q:=\sum_{i\in\mathbb{Z}}E_{i+1,i},

where Ei,jE_{i,j} is the matrix unit with 11 on the (i,j)(i,j)-entry, belong to 𝔤J(k)\mathfrak{g}J(k) but not to 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k). They satisfy

  1. 1.

    QtQ=I{}^{t}Q\cdot Q=I,

  2. 2.

    PQQP=IPQ-QP=I,

where II denotes the identity matrix iEi,i𝔤J(k)\sum_{i\in\mathbb{Z}}E_{i,i}\in\mathfrak{g}J(k). This matrix II does not even belong to 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k) ! (Off course, the matrices PP and QQ are matrix representations of ddt\dfrac{d}{dt} and tt\cdot on the vector space [t±1]\mathbb{C}[t^{\pm 1}] with respect to the basis {ti}i\{t^{i}\}_{i\in\mathbb{Z}}.)

Such algebras show up in many areas of mathematics and physics. For instance, they are used to describe the solitons of the Kadomtsev-Petviashvili (KP) type hierarchies [Sa] where such integrable systems are interpreted as a dynamical system on the so-called Sato Grassmannian. On the other hand, their basic algebraic properties and invariants are not well understood. In our work we present results on their homology with trivial coefficients, For this, we need to use several different (co)homology concepts. Some results were obtained by Feigin and Tsygan [FT] in 1983, but in their short note the statements and proofs are not precise. In our work, we were able to get straightforward statements and proofs and we also could generalize the results to the coefficients over an associative unital kk-algebra.

The structure of the paper is as follows. In Section 2 we introduce some important classes of Lie algebras of general Jacobi matrices and recall their universal central extension. We also give some examples of its subalgebras. Section 3 is devoted to their (co)homology. First we recall the main definitions: Lie algebra homology, Hochschild homology, cyclic homology and (skew)dihedral homology. Then we compute homology with trivial coefficients for the introduced Lie algebras. In this section, we also present precise proofs of some results in [FT] and give possible generalizations. In Section 4 we introduce two important subalgebras, the orthogonal and symplectic subalgebras. To compute their (co)homology, we need to introduce additional computational methods to the previous one. In Section 5 we discuss a more general case, where instead of the field kk we have an associative unital kk-algebra, and generalize our (co)homology results for such algebras. Finally, we introduce a rank functional on a Lie subalgebra 𝔤J(k)𝔤J(k)\mathfrak{g}J_{\infty}(k)\subseteq\mathfrak{g}J(k) and describe its image. After describing its cohomology ring, we also raise some open questions.

2 Lie algebras of generalized Jacobi matrices

2.1 Lie algebras 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k) and 𝔤J(k)\mathfrak{g}J(k)

The first such Lie algebra one may have in mind is the one defined as an inductive limit: let II be a countable set and I1I2InI_{1}\subsetneq I_{2}\subsetneq\cdots\subsetneq I_{n}\subsetneq\cdots, I=nInI=\bigcup_{n}I_{n} an increasing series where each InI_{n} is a finite subset. The Lie algebra 𝔤𝔩I(k)\mathfrak{gl}_{I}(k) is defined by the inductive limit of 𝔤𝔩Im(k)𝔤𝔩In(k)\mathfrak{gl}_{I_{m}}(k)\hookrightarrow\mathfrak{gl}_{I_{n}}(k) for m<nm<n, namely, each element of 𝔤𝔩I(k)\mathfrak{gl}_{I}(k) is a finitely supported matrix of infinite size indexed over II, i.e., X=(xi,j)i,jIX=(x_{i,j})_{i,j\in I} such that {(i,j)I2|xi,j0}<\sharp\{(i,j)\in I^{2}\,|\,x_{i,j}\neq 0\,\}<\infty. In particular, for I=I=\mathbb{Z}, we may denote by 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k).

Another one we may also encounter is the Lie algebra of generalized Jacobi matrices defined as follows. A generalized Jacobi matrix is a matrix M=(mi,j)M=(m_{i,j}) indexed over \mathbb{Z} such that there exists a positive integer NMN_{M} satisfying

mi,j=0i,jsuch that|ij|>NM.m_{i,j}=0\qquad\forall\;i,j\quad\text{such that}\quad|i-j|>N_{M}.

The set of such matrices has a structure of associative algebra over kk denoted by J(k)J(k). We shall denote it by 𝔤J(k)\mathfrak{g}J(k) whenever we regard it as Lie algebra. An original Jacobi matrix is a finite size matrix M=(mi,j)M=(m_{i,j}) such that mi,j=0m_{i,j}=0 for any i,ji,j with |ij|>1|i-j|>1. The Lie algebra 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k) can be naturally viewed as a subalgebra of 𝔤J(k)\mathfrak{g}J(k).

2.2 Universal central extension of the Lie algebra 𝔤J(k)\mathfrak{g}J(k)

In the course of studying soliton theory, a non-trivial central extension of the Lie algebra 𝔤J(k)\mathfrak{g}J(k) was discovered (see, e.g., [JM] and [DJM]) and it can be described as follows.

Let J=i0Ei,i𝔤J(k)J=\sum_{i\geq 0}E_{i,i}\in\mathfrak{g}J(k) be a matrix and let Φ:J(k)J(k)\Phi:J(k)\rightarrow J(k) be the kk-linear map defined by XJXJX\mapsto JXJ. It can be checked that, for any X,YJ(k)X,Y\in J(k), the element [Φ(X),Φ(Y)]Φ([X,Y])[\Phi(X),\Phi(Y)]-\Phi([X,Y]) is an element of 𝔤𝔩(k)\mathfrak{gl}_{\infty}(k), i.e., only finitely many matrix entires can be non-zero. Hence, one can define the kk-bilinear map Ψ:J(k)×J(k)k\Psi:J(k)\times J(k)\rightarrow k by

Ψ(X,Y)=tr([Φ(X),Φ(Y)]Φ([X,Y])),\Psi(X,Y)=\mathrm{tr}([\Phi(X),\Phi(Y)]-\Phi([X,Y])),

where tr:𝔤𝔩(k)k;X=(xi,j)ixi,i\mathrm{tr}:\mathfrak{gl}_{\infty}(k)\rightarrow k\,;\,X=(x_{i,j})\mapsto\sum_{i\in\mathbb{Z}}x_{i,i} is the trace of finitely supported matrices. It turned out that this Ψ\Psi is a 22-cocycle, called Japanese cocycle, i.e.,

  1. 1.

    Ψ(Y,X)=Ψ(X,Y)\Psi(Y,X)=-\Psi(X,Y),

  2. 2.

    Ψ([X,Y],Z)+Ψ([Y,Z],X)+Ψ([Z,X],Y)=0\Psi([X,Y],Z)+\Psi([Y,Z],X)+\Psi([Z,X],Y)=0,

for any X,Y,ZJ(k)X,Y,Z\in J(k). Let 𝔤J~(k):=𝔤J(k)k1\widetilde{\mathfrak{g}J}(k):=\mathfrak{g}J(k)\oplus k\cdot 1 be the Lie algebra whose Lie bracket [,][\cdot,\cdot]^{\prime} is given by

[X,1]=0,[X,Y]=[X,Y]+Ψ(X,Y)1X,Y𝔤J(k).[X,1]^{\prime}=0,\qquad[X,Y]^{\prime}=[X,Y]+\Psi(X,Y)1\qquad X,Y\in\mathfrak{g}J(k).

As the Lie algebra 𝔤J(k)\mathfrak{g}J(k) is perfect, i.e., [𝔤J(k),𝔤J(k)]=𝔤J(k)[\mathfrak{g}J(k),\mathfrak{g}J(k)]=\mathfrak{g}J(k), the Lie algebra 𝔤J(k)\mathfrak{g}J(k) admits the universal central extension. It was B. L. Feigin and B. L. Tsygan [FT] in 1983 who proved that this central extension α:𝔤J~(k)𝔤J(k)\alpha:\widetilde{\mathfrak{g}J}(k)\rightarrow\mathfrak{g}J(k) is universal, namely, 𝔤J~(k)\widetilde{\mathfrak{g}J}(k) is perfect and for any central extension β:𝔞𝔤J(k)\beta:\mathfrak{a}\rightarrow\mathfrak{g}J(k), there exists a morphism of Lie algebras γ:𝔤J~(k)𝔞\gamma:\widetilde{\mathfrak{g}J}(k)\rightarrow\mathfrak{a} such that the next diagram commutes:

𝔤J~(k)\textstyle{\widetilde{\mathfrak{g}J}(k)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}γ\scriptstyle{\gamma}𝔤J(k)\textstyle{\mathfrak{g}J(k)}𝔞\textstyle{\mathfrak{a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}
Remark 2.1.

The kernel 𝔷\mathfrak{z} of the universal central extension of 𝔤J(k)\mathfrak{g}J(k), i.e., the kernel of the canonical projection 𝔤J~(k)𝔤J(k)\widetilde{\mathfrak{g}J}(k)\twoheadrightarrow\mathfrak{g}J(k) can be given by the 22nd homology H2(𝔤J(k))H_{2}(\mathfrak{g}J(k)) and the 22-cocycle Ψ\Psi is an element of the 22nd cohomology H2(𝔤J(k))H^{2}(\mathfrak{g}J(k)).

Remark 2.2.

Let GJ~\widetilde{G_{J}} be the (pro-)algebraic group of 𝔤J~(k)\widetilde{\mathfrak{g}J}(k). The Lie algebra 𝔤J~(k)\widetilde{\mathfrak{g}J}(k) acts on a fermionic Fock space \mathcal{F}. The GJ~\widetilde{G_{J}}-orbit of its vacuum state |vac|\,\mathrm{vac}\rangle in the projective space \mathbb{P}\mathcal{F} is isomorphic to the Sato Grassmannian. The defining equations of this orbit in terms of Plücker coordinates is nothing but the Hirota bilinear equations of Kadomtsev-Petviashvili (KP) hierarchy. For details, see, e.g., [JM] and [DJM].

2.3 Some subalgebras of the universal central extension 𝔤J~(k)\widetilde{\mathfrak{g}J}(k)

The extended Lie algebra contains several interesting infinite dimensional Lie algebras as subalgebras. We shall introduce some of them.

1. For an integer n>1n>1, let 𝔤Jn~(k)\widetilde{\mathfrak{g}J_{n}}(k) be the subalgebra of 𝔤J~(k)\widetilde{\mathfrak{g}J}(k) generated by the matrices M=(mi,j)i,jM=(m_{i,j})_{i,j\in\mathbb{Z}} satisfying mi+n,j+n=mi,jm_{i+n,j+n}=m_{i,j} for any i,ji,j\in\mathbb{Z}. This subalgebra is isomorphic to the central extension of 𝔤𝔩n(k[t,t1])\mathfrak{gl}_{n}(k[t,t^{-1}]), i.e., the affine Lie algebra

𝔤𝔩n^(k)=𝔤𝔩n(k[t,t1])kc\widehat{\mathfrak{gl}_{n}}(k)=\mathfrak{gl}_{n}(k[t,t^{-1}])\oplus kc

whose commutation relation is given by

[ei,jta,ek,ltb]=[ei,j,ek,l]ta+b+aδa+b,0δj,kδi,lc,[𝔤𝔩n^(k),c]=0,[e_{i,j}\otimes t^{a},e_{k,l}\otimes t^{b}]=[e_{i,j},e_{k,l}]\otimes t^{a+b}+a\delta_{a+b,0}\delta_{j,k}\delta_{i,l}c,\qquad[\widehat{\mathfrak{gl}_{n}}(k),c]=0,

where ei,j𝔤𝔩ne_{i,j}\in\mathfrak{gl}_{n} is the matrix unit whose (i,j)(i,j)-entry is 11. An isomorphism between 𝔤𝔩n~(k)\widetilde{\mathfrak{gl}_{n}}(k) and 𝔤Jn~(k)\widetilde{\mathfrak{g}J_{n}}(k) is given by ei,jtarei+rn,j+(r+a)ne_{i,j}\otimes t^{a}\mapsto\sum_{r\in\mathbb{Z}}e_{i+rn,j+(r+a)n}.

2. Another important example is the one-dimensional central extension of the Lie algebra k[t,t1][ddt]k[t,t^{-1}]\left[\dfrac{d}{dt}\right] of algebraic differential operators over k=k{0}k^{\ast}=k\setminus\{0\} that is defined as follows. Set D=tddtD=t\frac{d}{dt}. For any polynomial f,gk[D]f,g\in k[D], it can be verified that

[trf(D),tsg(D)]=tr+s(f(D+s)g(D)f(D)g(D+r)).[t^{r}f(D),t^{s}g(D)]=t^{r+s}(f(D+s)g(D)-f(D)g(D+r)).

Let Ψ\Psi be the 22-cocyle on k[t,t1][ddt]k[t,t^{-1}]\left[\dfrac{d}{dt}\right] defined by

Ψ(trf(D),tsg(D)):={rj1f(j)g(j+r)r=s0,0r+s0.\Psi(t^{r}f(D),t^{s}g(D)):=\begin{cases}\sum_{-r\leq j\leq-1}f(j)g(j+r)\qquad&r=-s\geq 0,\\ \quad 0\quad&r+s\neq 0.\end{cases}

The Lie algebra 𝒲1+\mathcal{W}_{1+\infty} is, by definition, the central extension of k[t,t1][ddt]k[t,t^{-1}]\left[\dfrac{d}{dt}\right] by the 22-cocycle Ψ\Psi, i.e., it is the kk-vector space

𝒲1+=k[t,t1][ddt]kC,\mathcal{W}_{1+\infty}=k[t,t^{-1}]\left[\dfrac{d}{dt}\right]\oplus kC,

equipped with the Lie bracket given by

[trf(D),tsg(D)]=[trf(D),tsg(D)]+Ψ(trf(D),tsg(D))C,[𝒲1+,C]=0.[t^{r}f(D),t^{s}g(D)]^{\prime}=[t^{r}f(D),t^{s}g(D)]+\Psi(t^{r}f(D),t^{s}g(D))C,\qquad[\mathcal{W}_{1+\infty},C]=0.

It can be shown that there exists morphism of Lie algebras 𝒲1+𝔤J~(k)\mathcal{W}_{1+\infty}\hookrightarrow\widetilde{\mathfrak{g}J}(k) satisfying taDriiaEi+r,it^{a}D^{r}\,\mapsto\sum_{i\in\mathbb{Z}}i^{a}E_{i+r,i}. We remark that the Lie subalgebra k[t,t1]DkCk[t,t^{-1}]D\oplus kC of 𝒲1+\mathcal{W}_{1+\infty} is isomorphic to the Virasoro algebra.

The first example shows that the Lie algebra 𝔤J~(k)\widetilde{\mathfrak{g}J}(k) contains, at least, affine Lie algebras of classical type, i.e., Al(1)(l1),Bl(1)(l3),Cl(1)(l2),Dl(1)(l4),A2l(2)(l1),A2l1(2),(l3),Dl+1(2)(l2)A_{l}^{(1)}(l\geq 1),B_{l}^{(1)}(l\geq 3),C_{l}^{(1)}(l\geq 2),D_{l}^{(1)}(l\geq 4),A_{2l}^{(2)}(l\geq 1),A_{2l-1}^{(2)},(l\geq 3),D_{l+1}^{(2)}(l\geq 2) and D4(3)D_{4}^{(3)}. In addition, the second example shows that Lie algebra 𝔤J~(k)\widetilde{\mathfrak{g}J}(k) contains also the Lie algebra 𝒲1+\mathcal{W}_{1+\infty} that plays an important role in the KP-hierarchy.

3 Homology of the Lie algebra 𝔤J(k)\mathfrak{g}J(k)

In this Section, among others we state the main result of [FT] and explain the outline of the proof. Our may goal is the computation of the homology H(𝔤J(k))H_{\bullet}(\mathfrak{g}J(k)).

3.1 Several homologies

We briefly recall the definitions of Lie algebra (co)homology, Hochschild homology, cyclic homology and (skew-)dihedral homology.

Let 𝔤\mathfrak{g} be a Lie algebra over a field kk of characteristic 0. From now on, we shall abbreviate the coefficient kk. The Lie algebra homology H(𝔤)H_{\bullet}(\mathfrak{g}) is, by definition, the homology of the complex (𝔤,d)(\bigwedge^{\bullet}\mathfrak{g},d), called the Eilenberg-Chevalley complex, where 𝔤\bigwedge^{\bullet}\mathfrak{g} is the exterior algebra of 𝔤\mathfrak{g} and the differential dd is given by

d(x1xn)=1i<jn(1)i+j+1[xi,xj]x1xi^xj^xn.d(x_{1}\wedge\cdots\wedge x_{n})=\sum_{1\leq i<j\leq n}(-1)^{i+j+1}[x_{i},x_{j}]\wedge x_{1}\wedge\cdots\hat{x_{i}}\wedge\cdots\wedge\hat{x_{j}}\wedge\cdots\wedge x_{n}.

The Lie algebra cohomology H(𝔤)H^{\bullet}(\mathfrak{g}) is by definition, the cohomology of the ‘dual of the complex (𝔤,d)(\bigwedge^{\bullet}\mathfrak{g},d)’ , i.e., (Homk(𝔤,k),dt)(\mathrm{Hom}_{k}(\bigwedge^{\bullet}\mathfrak{g},k),-{}^{t}d). For some basic properties of this homology, see, e.g. [HS]. The homology H(𝔤)H_{\bullet}(\mathfrak{g}) has a commutative and cocommutative DG-Hopf algebra structure. Its coalgebra structure with counit is induced from the comultiplication

Δ:H(𝔤)𝛿H(𝔤𝔤)𝜎H(𝔤)H(𝔤),\Delta:H_{\bullet}(\mathfrak{g})\overset{\delta}{\longrightarrow}H_{\bullet}(\mathfrak{g}\oplus\mathfrak{g})\overset{\sigma}{\longrightarrow}H_{\bullet}(\mathfrak{g})\otimes H_{\bullet}(\mathfrak{g}),

where δ:𝔤𝔤𝔤;x(x,x)\delta:\mathfrak{g}\rightarrow\mathfrak{g}\oplus\mathfrak{g};\,x\,\mapsto(x,x) is the diagonal map, and σ\sigma is the Künneth isomorphism. Hence, the homology H(𝔤)H_{\bullet}(\mathfrak{g}) is the graded symmetric algebra over its primitive part (cf. [Q]). Recall that an element xH(𝔤)x\in H_{\bullet}(\mathfrak{g}) is said to be primitive if its cocommutative coalgebra structure with counit is induced from the comultiplication which satisfies Δ(x)=x1+1x\Delta(x)=x\otimes 1+1\otimes x. For the Lie algebra 𝔤𝔩(R)\mathfrak{gl}_{\infty}(R) over an associative unital kk-algebra RR, the primitive part of the homology H(𝔤𝔩(R))H_{\bullet}(\mathfrak{gl}_{\infty}(R)) had been known by Loday and Quillen [LQ] and independently by B. L. Tsygan [Ts] as follows:

Theorem 3.1.

The primitive part PrimH(𝔤𝔩(R))\mathrm{Prim}\,H_{\bullet}(\mathfrak{gl}_{\infty}(R)) is isomorphic to the cyclic homology HC1(R)HC_{\bullet-1}(R).

The cohomology ring H(𝔤)H^{\bullet}(\mathfrak{g}) is naturally endowed with a commutative and cocommutative DG-Hopf algebra structure. Indeed, by the Poincaré duality Hn(𝔤)Hn(𝔤)H^{n}(\mathfrak{g})\cong H_{n}(\mathfrak{g})^{\ast} (full dual), one has H(𝔤)H(𝔤):=nHn(𝔤)H^{\bullet}(\mathfrak{g})\cong H_{\bullet}(\mathfrak{g})^{\vee}:=\bigoplus_{n}H_{n}(\mathfrak{g})^{\ast} (restricted dual).

Now we recall some definitions in homology theory of associative algebras. For detail, see, e.g., [Lod].

Let RR be an associative unital kk-algebra and MM be an RR-bimodule. For n0n\in\mathbb{Z}_{\geq 0}, set Cn(R,M)=MRnC_{n}(R,M)=M\otimes R^{\otimes n}. The Hochschild homology H(R,M)H_{\bullet}(R,M) is, by definition, the homology of the complex (C(R,M),b)(C_{\bullet}(R,M),b), where the the differential bb is defined by

b(mr1rn)\displaystyle b(m\otimes r_{1}\otimes\cdots\otimes r_{n})
=\displaystyle= ma1a2an+i=1n1(1)ima1aiai+1an\displaystyle ma_{1}\otimes a_{2}\otimes\cdots\otimes a_{n}+\sum_{i=1}^{n-1}(-1)^{i}m\otimes a_{1}\otimes\cdots\otimes a_{i}a_{i+1}\otimes\cdots\otimes a_{n}
+(1)nanma1an1.\displaystyle+(-1)^{n}a_{n}m\otimes a_{1}\otimes\cdots\otimes a_{n-1}.

In particular, for M=RM=R, for each n>0n\in\mathbb{Z}_{>0}, there is an action of the cyclic group /(n+1)\mathbb{Z}/(n+1)\mathbb{Z} given by

x.(r0r1rn)=(1)n(rnr0rn1),x.(r_{0}\otimes r_{1}\otimes\cdots\otimes r_{n})=(-1)^{n}(r_{n}\otimes r_{0}\otimes\cdots\otimes r_{n-1}),

where xx is a generator of the group /(n+1)\mathbb{Z}/(n+1)\mathbb{Z}. The differential bb of the complex (C(R,R),b)(C_{\bullet}(R,R),b) induces a differential on the complex Cnλ(R):=R(n+1)/(1x)C_{n}^{\lambda}(R):=R^{\otimes(n+1)}/(1-x). The homology of this complex, called Connes’ complex is the so-called cyclic homology HC(R)HC_{\bullet}(R) of RR. For some cases, this cyclic homology can be computed with the aid of Connes’ periodicity exact sequence:

HHn(R)HCn(R)HCn2(R)HHn1(R).\cdots\rightarrow HH_{n}(R)\rightarrow HC_{n}(R)\rightarrow HC_{n-2}(R)\rightarrow HH_{n-1}(R)\rightarrow\cdots. (1)

Now, assume that RR is equipped with a kk-linear anti-involution ¯:RR\bar{\cdot}:R\rightarrow R. One can extend the action of the group /(n+1)\mathbb{Z}/(n+1)\mathbb{Z} on R(n+1)R^{\otimes(n+1)} to the dihedral group Dn+1=x,y|xn+1=y2=1,yxy=x1D_{n+1}=\langle x,y|x^{n+1}=y^{2}=1,yxy=x^{-1}\rangle by

y.(r0r1rn)=(1)12n(n+1)(r0¯rn¯rn1¯r1¯).y.(r_{0}\otimes r_{1}\otimes\cdots\otimes r_{n})=(-1)^{\frac{1}{2}n(n+1)}(\overline{r_{0}}\otimes\overline{r_{n}}\otimes\overline{r_{n-1}}\otimes\cdots\otimes\overline{r_{1}}).

Let 𝐃n(R)\mathbf{D}_{n}(R) denote the space of coinvariants (Rn+1)Dn+1(R^{\otimes n+1})_{D_{n+1}}. If we modify the action of yy by y-y, the resulting coinvariants will be denoted by 𝐃n1(R){}_{-1}\mathbf{D}_{n}(R). The differential bb on C(R,R)C_{\bullet}(R,R) induces the differentials on (Rn+1)Dn+1(R^{\otimes n+1})_{D_{n+1}} and 𝐃n1(R){}_{-1}\mathbf{D}_{n}(R), denoted by b¯\bar{b}. Their homologies are called dihedral (resp. skew-dihedral) homology of RR:

HDn(R):=Hn(𝐃(R),b¯),(resp.H1Dn(R):=Hn(𝐃1(R),b¯)).HD_{n}(R):=H_{n}(\mathbf{D}_{\bullet}(R),\bar{b}),\qquad(\;\text{resp}.\;{}_{-1}HD_{n}(R):=H_{n}({}_{-1}\mathbf{D}_{\bullet}(R),\bar{b})\;).

For more informations, see, e.g., [Lod].

3.2 An isomorphism 𝔤J(k)𝔤𝔩n(J(k))\mathfrak{g}J(k)\cong\mathfrak{gl}_{n}(J(k))

Let n>1n>1 be an integer. We fix a section of /n\mathbb{Z}\twoheadrightarrow\mathbb{Z}/n\mathbb{Z} and denote its image by II. For any matrix M=(mi,j)i,jM=(m_{i,j})_{i,j\in\mathbb{Z}} and any i,jIi,j\in I, we set Mi,j:=(mi+nk,j+nl)k,lM_{i,j}:=(m_{i+nk,j+nl})_{k,l\in\mathbb{Z}}. Then the kk-linear map

ΦI:𝔤J(k)𝔤𝔩n(J(k));M(Mi,j)i,jI,\Phi_{I}:\mathfrak{g}J(k)\longrightarrow\mathfrak{gl}_{n}(J(k));\qquad M\;\longmapsto\;(M_{i,j})_{i,j\in I},

is an isomorphism of Lie algebras 𝔤J(k)𝔤𝔩n(J(k))\mathfrak{g}J(k)\longrightarrow\mathfrak{gl}_{n}(J(k)), and it induces an isomorphism of homologies H(𝔤J(k))H(𝔤𝔩n(J(k)))H_{\bullet}(\mathfrak{g}J(k))\cong H_{\bullet}(\mathfrak{gl}_{n}(J(k))). Taking an inductive limit, we obtain

Lemma 3.2 (Lemma 1 of [FT]).

H(𝔤J(k))H(𝔤𝔩(J(k)))H_{\bullet}(\mathfrak{g}J(k))\cong H_{\bullet}(\mathfrak{gl}_{\infty}(J(k))).

Hence, the homology H(𝔤J(k))H_{\bullet}(\mathfrak{g}J(k)) is the commutative and cocommutative DG-Hopf algebra whose primitive part is HC1(J(k))HC_{\bullet-1}(J(k)) by Theorem 3.1. Thus, if we can compute the Hochschild homology HH(J(k))HH_{\bullet}(J(k)) of the associative algebra J(k)J(k), Connes’ periodicity exact sequence (1) allows us to determine the cyclic homology HC(J(k))HC_{\bullet}(J(k)).

In the rest of this section, we will explain this briefly.

3.3 Twisted Group Algebras

Let AA be an associative unital kk-algebra and let GG be a discrete subgroup of the group of kk-automorphisms of AA. One can twist the natural product structure on Akk[G]A\otimes_{k}k[G] by (a[g])(b[h]):=ag(b)[gh](a\otimes[g])\cdot(b\otimes[h]):=ag(b)\otimes[gh], where a,bAa,b\in A and g,hGg,h\in G. The tensor product Akk[G]A\otimes_{k}k[G] equipped with such a twisted product is called twisted group algebra and will be denoted by A{G}A\{G\}.

Example 3.3.
  1. 1.

    Fix n>1n\in\mathbb{Z}_{>1}. Let A=k×k××knA=\overbrace{k\times k\times\cdots\times k}^{n} be the nn-copies of kk viewed as commutative associative algebra. The group G=/nG=\mathbb{Z}/n\mathbb{Z} acts on A via cyclic permutation: (i+n).(a1,,an)=(ai+1,,ai+n)(i+n\mathbb{Z}).(a_{1},\ldots,a_{n})=(a_{i+1},\ldots,a_{i+n}) where the indicies should be understood modulo nn. Then, the twisted group algebra A{G}A\{G\} is isomorphic to the algebra of n×nn\times n matrices Mn(k)M_{n}(k). The isomorphism A{G}Mn(k)A\{G\}\rightarrow M_{n}(k) is given by

    (a1,,an)[i+n]k=1nakek,k+i.(a_{1},\ldots,a_{n})\otimes[i+n\mathbb{Z}]\quad\longmapsto\quad\sum_{k=1}^{n}a_{k}e_{k,k+i}.
  2. 2.

    A=ikeiA=\prod_{i\in\mathbb{Z}}ke_{i}, where eie_{i}’s are orthogonal idempotents and G=G=\mathbb{Z}. The additive group \mathbb{Z} acts on AA: 1.ei=ei1.1.e_{i}=e_{i-1}. One can show that the twisted group algebra A{G}A\{G\} is isomorphic to the associative algebra J(k)J(k) where the isomorphism A{G}J(k)A\{G\}\rightarrow J(k) is given by

    kakek[i]kakEk,k+i.\sum_{k\in\mathbb{Z}}a_{k}e_{k}\otimes[i]\quad\longmapsto\quad\sum_{k\in\mathbb{Z}}a_{k}E_{k,k+i}.

3.4 Hochschild-Serre type Spectral Sequence

Let MM be an A{G}A\{G\}-bimodule.

Theorem 3.4.

[St] There exists a spectral sequence

Ep,q2=Hp(k[G],Hq(A,M))Hp+q(A{G},M)E_{p,q}^{2}=H_{p}(k[G],H_{q}(A,M))\;\Longrightarrow\;H_{p+q}(A\{G\},M)
Remark 3.5.

In [FT], B. Feigin and B. Tsygan treated the above case when M=A{G}M=A\{G\}. But their description of the k[G]k[G]-module structure on Hq(A,A{G})H_{q}(A,A\{G\}) was not correct and this was rectified by D. Stefan [St] under more general setting.

In our case, G=G=\mathbb{Z} and k[]k[\mathbb{Z}] is a principal ideal domain, hence its global dimension is at most 11 which implies that this spectral sequence collapses at E2E^{2}.

By a standard argument, one can show that

Hq(A,J(k)))iHHq(k)ei.H_{q}(A,J(k)))\cong\prod_{i\in\mathbb{Z}}HH_{q}(k)e_{i}.

Moreover, by direct computation, it can be verified that

Hp(k[],Hq(A,J(k))){HHq(k)p=1,0p1.H_{p}(k[\mathbb{Z}],H_{q}(A,J(k)))\cong\begin{cases}HH_{q}(k)\qquad&p=1,\\ \quad 0\quad&p\neq 1.\end{cases}

This implies HHp(J(k))HHp1(k).HH_{p}(J(k))\cong HH_{p-1}(k).

Remark 3.6.

The isomorphism H1(k[],Hp(A,J(k))HHp+1(J(k))H_{1}(k[\mathbb{Z}],H_{p}(A,J(k))\cong HH_{p+1}(J(k)) is given by the so-called shuffle product [Q]. This fact plays an important role when we determined the homology of the Lie algebras of orthogonal and symplectic generalized Jacobi matrices in [FI2].

By definition, it follows that

HHp(k)={kp=0,0p>0.HH_{p}(k)=\begin{cases}\quad k\quad&p=0,\\ \quad 0\quad&p>0.\end{cases}

Thus, we obtain

Theorem 3.7 (cf. Theorem 3 in [FT]).

HH1(J(k))=kHH_{1}(J(k))=k and HHp(J(k))=0HH_{p}(J(k))=0 for any p1p\neq 1.

By Connes’ Periodicity long exact sequence (1), one has

Corollary 3.8.

HCp(J(k))=kHC_{p}(J(k))=k for odd pp and HCp(J(k))=0HC_{p}(J(k))=0 for even pp.

3.5 Description of the primitive part Prim(H(𝔤J(k)))\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}J(k)))

Theorem 3.1 and Corollary 3.8 imply

Theorem 3.9.

Prim(H(𝔤J(k)))\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}J(k))) is the graded kk-vector space whose ppth graded component is

Prim(H(𝔤J(k)))p={kp0(2), 0otherwise.\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}J(k)))_{p}=\begin{cases}\,k\,\quad&p\equiv 0(2),\\ \,0\,&\text{otherwise}.\end{cases}

In particular, each graded subspace of H(𝔤J(k))H_{\bullet}(\mathfrak{g}J(k)) is of finite dimension. Thus, by the Poincaré duality Hp(𝔤J(k))Hp(𝔤J(k))H^{p}(\mathfrak{g}J(k))\cong H_{p}(\mathfrak{g}J(k))^{\ast}, one obtains Theorem 1. a) in [FT], i.e., the cohomology ring H(𝔤J(k))H^{\bullet}(\mathfrak{g}J(k)) has the next description:

Theorem 3.10.

There exists primitive elements ciH2i(𝔤J(k))c_{i}\in H^{2i}(\mathfrak{g}J(k)) such that H(𝔤J(k))H^{\bullet}(\mathfrak{g}J(k)) is isomorphic to the Hopf algebra S(i>0kci)S(\bigoplus_{i\in\mathbb{Z}_{>0}}kc_{i}).

4 Orthogonal and Symplectic Subalgebras of 𝔤J(k)\mathfrak{g}J(k)

In this Section, after recalling the definition of orthogonal and symplectic subalgebras of 𝔤J(k)\mathfrak{g}J(k), we present two key steps to compute their homology of these subalgebras and state the results about their homology.

4.1 Definitions

Let RR be an associative unital kk-algebra equipped with an kk-linear anti-involution ¯\bar{\cdot}, i.e., s¯¯=s\bar{\bar{s}}=s and st¯=t¯s¯\overline{st}=\bar{t}\bar{s} for s,tRs,t\in R. We extend the transpose, also denoted by ()t{}^{t}(\cdot), to J(R)J(R) by (Ei,j(r))t=Ej,i(r¯){}^{t}(E_{i,j}(r))=E_{j,i}(\bar{r}), where Ei,j(r)=Ei,jrJ(R)E_{i,j}(r)=E_{i,j}r\in J(R).

For ll\in\mathbb{Z}, let τl,τls\tau_{l},\tau_{l}^{s} be the kk-linear anti-involutions of the Lie algebra 𝔤J(k)\mathfrak{g}J(k) defined by

τl(X)=(1)lJl(Xt)Jl,Jl:=i(1)iEi,li,\displaystyle\tau_{l}(X)=(-1)^{l}J_{l}({}^{t}X)J_{l},\qquad J_{l}:=\sum_{i\in\mathbb{Z}}(-1)^{i}E_{i,l-i},
τls(X)=Jls(Xt)Jls,Jls:=lEi,li.\displaystyle\tau_{l}^{s}(X)=J_{l}^{s}({}^{t}X)J_{l}^{s},\qquad J_{l}^{s}:=\sum_{l\in\mathbb{Z}}E_{i,l-i}.

We set

𝔬Jodd(k)=\displaystyle\mathfrak{o}_{J}^{\mathrm{odd}\,}(k)= 𝔤J(k)τ0s,={X𝔤J(k)|τ0s(X)=X},\displaystyle\mathfrak{g}J(k)^{\tau_{0}^{s},-}=\{X\in\mathfrak{g}J(k)|\tau_{0}^{s}(X)=-X\},
𝔰𝔭J(k)=\displaystyle\mathfrak{sp}_{J}(k)= 𝔤J(k)τ1,={X𝔤J(k)|τ1(X)=X},\displaystyle\mathfrak{g}J(k)^{\tau_{-1},-}=\{X\in\mathfrak{g}J(k)|\tau_{-1}(X)=-X\},
𝔬Jeven(k)=\displaystyle\mathfrak{o}_{J}^{\mathrm{even}\,}(k)= 𝔤J(k)τ1s,={X𝔤J(k)|τ1s(X)=X}.\displaystyle\mathfrak{g}J(k)^{\tau_{-1}^{s},-}=\{X\in\mathfrak{g}J(k)|\tau_{-1}^{s}(X)=-X\}.
Remark 4.1.

The universal central extension of the Lie algebras 𝔤J(k),𝔬Jodd(k),𝔰𝔭J(k)\mathfrak{g}J(k),\mathfrak{o}_{J}^{\mathrm{odd}\,}(k),\mathfrak{sp}_{J}(k) and 𝔬Jeven(k)\mathfrak{o}_{J}^{\mathrm{even}\,}(k) are the Lie algebras of type AJ,BJ,CJA_{J},B_{J},C_{J} and DJD_{J} respectively that are used to obtain the Hirota bilinear forms with these symmetries. See, e.g., [JM], for details.

4.2 Stable limit of Orthogonal and Symplectic Lie algebras and homology

Set

JB=iei,i,JC=i(1)iei,i1,JD=iei,i1.J_{B}=\sum_{i\in\mathbb{Z}}e_{i,-i},\qquad J_{C}=\sum_{i\in\mathbb{Z}}(-1)^{i}e_{i,-i-1},\qquad J_{D}=\sum_{i\in\mathbb{Z}}e_{i,-i-1}.

We define the anti-involutions τB,τC\tau_{B},\tau_{C} and τD\tau_{D} of 𝔤𝔩(R)\mathfrak{gl}_{\infty}(R) as follows:

τB(X)=JB(Xt)JB,τC(X)=JC(Xt)JC,τD(X)=JD(Xt)JD.\tau_{B}(X)=J_{B}({}^{t}X)J_{B},\qquad\tau_{C}(X)=-J_{C}({}^{t}X)J_{C},\qquad\tau_{D}(X)=J_{D}({}^{t}X)J_{D}.

The Lie algebras

𝔬odd(R):=𝔤𝔩(R)τB,,𝔰𝔭(R):=𝔤𝔩(R)τC,,𝔬even(R):=𝔤𝔩(R)τD,\mathfrak{o}_{\mathrm{odd}\,}(R):=\mathfrak{gl}_{\infty}(R)^{\tau_{B},-},\qquad\mathfrak{sp}(R):=\mathfrak{gl}_{\infty}(R)^{\tau_{C},-},\qquad\mathfrak{o}_{\mathrm{even}\,}(R):=\mathfrak{gl}_{\infty}(R)^{\tau_{D},-}

are clearly the stable limit of the Lie algebras {𝔬2l+1(R)}l>0,{𝔰𝔭2l(R)},{𝔬2l(R)}l>1\{\mathfrak{o}_{2l+1}(R)\}_{l\in\mathbb{Z}_{>0}},\{\mathfrak{sp}_{2l}(R)\},\{\mathfrak{o}_{2l}(R)\}_{l\in\mathbb{Z}_{>1}}, respectively. In a way similar to the case 𝔤J(k)\mathfrak{g}J(k) (cf. Lemma 3.2), one can show

Lemma 4.2.

There exists isomorphisms of homologies:

  1. 1.

    H(𝔬Jodd(k))H(𝔬odd(J(k)))H_{\bullet}(\mathfrak{o}^{\mathrm{odd}\,}_{J}(k))\cong H_{\bullet}(\mathfrak{o}_{\mathrm{odd}\,}(J(k))),

  2. 2.

    H(𝔰𝔭J(k))H(𝔰𝔭(J(k)))H_{\bullet}(\mathfrak{sp}_{J}(k))\cong H_{\bullet}(\mathfrak{sp}(J(k))),

  3. 3.

    H(𝔬Jeven(k))H(𝔬even(J(k)))H_{\bullet}(\mathfrak{o}^{\mathrm{even}\,}_{J}(k))\cong H_{\bullet}(\mathfrak{o}_{\mathrm{even}\,}(J(k))),

Thanks to Theorem 5.5 in [LP], due to J. L. Loday and C. Procesi, we obtain

Theorem 4.3.

Let :J(k)J(k)\ast:J(k)\rightarrow J(k) be the anti-involution satisfying

Er,s(a)=Es,r(a)ak.E_{r,s}(a)^{\ast}=E_{-s,-r}(a)\qquad a\in k.

Then, for 𝔤=𝔬Jodd,𝔰𝔭J\mathfrak{g}=\mathfrak{o}_{J}^{\mathrm{odd}\,},\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}, we have

Prim(H(𝔤(k)))=H1D1(J(k)),\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k)))={}_{-1}HD_{\bullet-1}(J(k)),

where H1D(){}_{-1}HD_{\bullet}(\cdot) signifies the skew-dihedral homology.

It is slightly technical, but one can show (cf. [FI2]) that

H1D1(J(k))=HD2(k).{}_{-1}HD_{\bullet-1}(J(k))=HD_{\bullet-2}(k).

Notice that HDp(k)=kHD_{p}(k)=k for p0mod 4p\equiv 0\,\text{mod}\,4 and HDp(k)=0HD_{p}(k)=0 otherwise. Thus, we have

Theorem 4.4 (cf. [FI2]).

For 𝔤=𝔬Jodd,𝔰𝔭J\mathfrak{g}=\mathfrak{o}_{J}^{\mathrm{odd}\,},\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}, Prim(H(𝔤(k)))\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k))) is the graded kk-vector space whose ppth graded component is

Prim(H(𝔤(k)))p={kp2(4), 0otherwise.\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k)))_{p}=\begin{cases}\,k\,\quad&p\equiv 2(4),\\ \,0\,\quad&\text{otherwise}.\end{cases}
Remark 4.5.

By Theorems 3.9 and 4.4, the support of Prim(H(𝔤(k)))\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k))) for 𝔤=𝔤J,𝔬Jodd,\mathfrak{g}=\,\mathfrak{g}J,\,\mathfrak{o}_{J}^{\mathrm{odd}\,},𝔰𝔭J\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}, i.e., those integers pp where Prim(H(𝔤(k)))p0\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k)))_{p}\neq 0, are exactly the twice of the exponents of the Weyl group of type AJ,BJ,CJA_{J},B_{J},C_{J} and DJD_{J}, respectively.

5 Generalizations and some further topics

Here, we give a generalization on the coefficient ring, mention some other related topics and open questions.

5.1 Lie algebras with coefficients in RR

The computations of Lie algebra homology given in Sections 2 and 3 can be generalized to the Lie algebra 𝔤(R)\mathfrak{g}(R), where 𝔤=𝔤J,𝔬Jodd,𝔰𝔭J\mathfrak{g}=\mathfrak{g}J,\mathfrak{o}_{J}^{\mathrm{odd}\,},\,\,\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}. Here RR is an associative unital kk-algebra (equipped with a kk-linear anti-involution ¯:RR\bar{\cdot}:R\rightarrow R for 𝔤=𝔬Jodd,𝔰𝔭J\mathfrak{g}=\mathfrak{o}_{J}^{\mathrm{odd}\,},\,\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}).

The next theorem generalizes Theorem 3.9:

Theorem 5.1 (cf. [FI1]).
  1. 1.

    Prim(H(𝔤J(R)))=HC2(R)\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}J(R)))=HC_{\bullet-2}(R),

  2. 2.

    H(𝔤J(R))S(HC2(R))H_{\bullet}(\mathfrak{g}J(R))\cong S(HC_{\bullet-2}(R)).

The proof goes as follows. As in the case R=kR=k, By Theorem 3.1, the primitive part Prim(H(𝔤J(R)))\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}J(R))) is given by HC1(J(R))HC_{\bullet-1}(J(R)). Now, D. Stefan’s spectral sequence allows us to show an isomorphism HH(J(R))HH1(R)HH_{\bullet}(J(R))\cong HH_{\bullet-1}(R). Analyzing carefully this isomorphism, it can be shown that HC1(J(R))HC2(R)HC_{\bullet-1}(J(R))\cong HC_{\bullet-2}(R).

The above proof also allows us to obtain a generalization of Theorem 4.4:

Theorem 5.2 (cf. [FI2]).

Set 𝔤=𝔬Jodd,𝔰𝔭J\mathfrak{g}=\mathfrak{o}_{J}^{\mathrm{odd}\,},\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}.

  1. 1.

    Prim(H(𝔤(k)))=HD2(R)\mathrm{Prim}\,(H_{\bullet}(\mathfrak{g}(k)))=HD_{\bullet-2}(R),

  2. 2.

    H(𝔤(R))S(HD2(R))H_{\bullet}(\mathfrak{g}(R))\cong S(HD_{\bullet-2}(R)).

5.2 Rank and trace functionals

For any n>1n\in\mathbb{Z}_{>1}, let 𝔤Jn(k)\mathfrak{g}J_{n}(k) be the image of the composition 𝔤Jn~(k)𝔤J~(k)𝔤J(k)\widetilde{\mathfrak{g}J_{n}}(k)\hookrightarrow\widetilde{\mathfrak{g}J}(k)\twoheadrightarrow\mathfrak{g}J(k). Let 𝔤J(k)\mathfrak{g}J_{\infty}(k) be the subalgebra of 𝔤J(k)\mathfrak{g}J(k) generated by all of the 𝔤Jn(k)\mathfrak{g}J_{n}(k) (n>1n\in\mathbb{Z}_{>1}). Viewed as an associative algebra, 𝔤Jn(k)\mathfrak{g}J_{n}(k) and 𝔤J(k)\mathfrak{g}J_{\infty}(k) will be denoted by Jn(k)J_{n}(k) and J(k)J_{\infty}(k), respectively. We don’t know whether this algebra J(k)J_{\infty}(k) is a von Neumann regular ring, i.e., for any AJ(k)A\in J_{\infty}(k), there exists XJ(k)X\in J_{\infty}(k) such that AXA=AAXA=A. Nevertheless, there is a so-called rank functional defined as follows. For A=(ai,j)i,jJ(k)A=(a_{i,j})_{i,j\in\mathbb{Z}}\in J_{\infty}(k), set

Rank(A)=limn12n+1rank(An),\mathrm{Rank}(A)=\lim_{n\rightarrow\infty}\frac{1}{2n+1}\mathrm{rank}(A_{n}),

where, the matrix AnA_{n} of size 2n+12n+1 is defined by (ai,j)ni,jn(a_{i,j})_{-n\leq i,j\leq n} and rank()\mathrm{rank}(\,\cdot\,) signifies the rank of a finite size matrix.

Remark 5.3.

It was pointed out by B. Feigin and B. Tsygan in [FT] that one can define the trace functional Tr:J(k)k\mathrm{Tr}:J_{\infty}(k)\rightarrow k by

Tr(A)=limn12n+1tr(An),\mathrm{Tr}(A)=\lim_{n\rightarrow\infty}\frac{1}{2n+1}\mathrm{tr}(A_{n}),

where tr()\mathrm{tr}(\cdot) is the trace of a finite size matrix. They used this functional to describe non trivial cocycles of the cohomology H(𝔤J(k))H^{\bullet}(\mathfrak{g}J_{\infty}(k)).

It turns out that Im(Rank)=[0,1]\mathrm{Im}(\mathrm{Rank})=[0,1]\cap\mathbb{Q}. Indeed, this rank functional is a rank function on J(k)J_{\infty}(k) in the sense of J. von Neumann [vN], i.e., it satisfies,

  1. 1.

    Rank(I)=1\mathrm{Rank}(I)=1,

  2. 2.

    Rank(xy)Rank(x),Rank(y)\mathrm{Rank}(xy)\leq\mathrm{Rank}(x),\mathrm{Rank}(y),

  3. 3.

    Rank(e+f)=Rank(e)+Rank(f)\mathrm{Rank}(e+f)=\mathrm{Rank}(e)+\mathrm{Rank}(f) for all orthogonal idempotents e,fJ(k)e,f\in J_{\infty}(k), and

  4. 4.

    Rank(x)=0\mathrm{Rank}(x)=0 if and only if x=0x=0.

Hence, let J^(k)J(k)\widehat{J_{\infty}}(k)\subset J(k) be the Rank\mathrm{Rank}-completion of J(k)J_{\infty}(k). We denote the continuous extension of Rank()\mathrm{Rank}(\,\cdot\,) to J^(k)\widehat{J_{\infty}}(k) by Rank¯()\overline{\mathrm{Rank}}(\,\cdot\,) (cf. [G]).

Proposition 5.4.

Im(Rank¯)=[0,1]\mathrm{Im}(\overline{\mathrm{Rank}})=[0,1].

Proof.

It suffices to show that for any x[0,1]cx\in[0,1]\cap\mathbb{Q}^{c}, there exists a diagonal D=diag(di)iD=\mathrm{diag}(d_{i})_{i\in\mathbb{Z}} such that Rank¯(D)=x\overline{\mathrm{Rank}}(D)=x.
Let {rn}n0>0\{r_{n}\}_{n\in\mathbb{Z}_{\geq 0}}\subset\mathbb{Z}_{>0} be the sequence defined as follows. Set r0=1r_{0}=1. For n>0n>0, let rnr_{n} be the integer such that rn12n+1<x<rn2n+1\frac{r_{n}-1}{2n+1}<x<\frac{r_{n}}{2n+1}. Such an integer exists since xx is irrational. It can be checked that if 0<x<120<x<\frac{1}{2}, then rn+1rn{0,1}r_{n+1}-r_{n}\in\{0,1\}, otherwise, rn+1rn{1,2}r_{n+1}-r_{n}\in\{1,2\}.
Now, we define the sequence {di}i\{d_{i}\}_{i\in\mathbb{Z}}. Set d0=1d_{0}=1. Suppose that {di}|i|n\{d_{i}\}_{|i|\leq n} is defined. Choose d±(n+1){0,1}d_{\pm(n+1)}\in\{0,1\} in such a way that {i{±(n+1)}|di=1}=rn+1rn\sharp\{i\in\{\pm(n+1)\}|d_{i}=1\}=r_{n+1}-r_{n}. For any such choice, it follows that the rank of the diagonal matrix diag(di)|i|n\mathrm{diag}(d_{i})_{|i|\leq n} is rnr_{n} for any n0n\in\mathbb{Z}_{\geq 0}. Hence, we have Rank¯(D)=x\overline{\mathrm{Rank}}(D)=x by construction. ∎

There might be a von Neumann regular subalgebra of J(k)J(k) that has not been explored up to now.

Here are some questions:

  1. 1.

    Does J^(k)\widehat{J_{\infty}}(k) admit a Lie algebra structure ? If yes,

    1. i)

      and if the Lie algebra 𝔤J^(k)\mathfrak{g}\widehat{J_{\infty}}(k) is perfect, describe its universal central extension.

    2. ii)

      More generally, can we have a description of the (co)homology ring of Lie algebra 𝔤J^(k)\mathfrak{g}\widehat{J_{\infty}}(k) ?

  2. 2.

    Is J(k)J_{\infty}(k) von Neumann regular algebra ? If yes, so is J^(k)\widehat{J_{\infty}}(k) (cf. [G]).

    1. i)

      What kind of additional properties, do they have ?

5.3 More about 𝔤J(k)\mathfrak{g}J_{\infty}(k)

In [FT], B. Feigin and B. Tsygan determined the structure of the cohomology ring H(𝔤J(k))H^{\bullet}(\mathfrak{g}J(k)) which states, for each i>0i>0, there exists, up to scalar, unique ciH2i(𝔤J(k))c_{i}\in H^{2i}(\mathfrak{g}J(k)) such that

H(𝔤J(k))S(i>0kci).H^{\bullet}(\mathfrak{g}J(k))\cong S^{\bullet}(\bigoplus_{i>0}kc_{i}).

This follows from Theorem 3.9. The natural inclusion 𝔤J(k)𝔤J(k)\mathfrak{g}J_{\infty}(k)\hookrightarrow\mathfrak{g}J(k) induces a surjective map H(𝔤J(k))H(𝔤J(k))H^{\bullet}(\mathfrak{g}J_{\infty}(k))\twoheadrightarrow H^{\bullet}(\mathfrak{g}J(k)). Indeed, the showed that, for each i>0i>0, there exists, up to scalar, unique ξiH2i1(𝔤J(k))\xi_{i}\in H^{2i-1}(\mathfrak{g}J_{\infty}(k)) such that the next short sequence is exact:

0(i>0kξi)H(𝔤J(k))H(𝔤J(k))0.0\longrightarrow\bigwedge{}^{\bullet}(\bigoplus_{i>0}k\xi_{i})\longrightarrow H^{\bullet}(\mathfrak{g}J_{\infty}(k))\longrightarrow H^{\bullet}(\mathfrak{g}J(k))\longrightarrow 0.

They even presented an explicit realization of cocycles.

It may be an interesting problem to obtain such descriptions also for the Lie algebras 𝔤J(k):=𝔤J(k)𝔤J(k)\mathfrak{g}_{J_{\infty}}(k):=\mathfrak{g}_{J}(k)\cap\mathfrak{g}J_{\infty}(k), where 𝔤J=𝔬Jodd,𝔰𝔭J\mathfrak{g}_{J}=\mathfrak{o}_{J}^{\mathrm{odd}\,},\mathfrak{sp}_{J} and 𝔬Jeven\mathfrak{o}_{J}^{\mathrm{even}\,}.

Acknowledgments

The authors thank Max Karoubi and Andrey Lazarev for helpful discussions. Kenji Iohara would like to thank the organizers for giving him the opportunity to present some of the results in this article.

References

  • [B] O. Bräunling, On the homology of Lie algebras like 𝔤𝔩(,R)\mathfrak{gl}(\infty,R), Homology, Homotopy and Appl. 21 (2019), 131–143.
  • [DJM] E. Date, M. Jimbo and T. Miwa, Solitons: Differential Equations, Symmetries and Infinite Dimensional Algebras ((translated from the original Japanese version by M. Reid)), Cambridge tracts in Math. 135, 2012.
  • [FI1] A. Fialowski and K. Iohara Homology of the Lie algebra 𝔤𝔩(,R)\mathfrak{gl}(\infty,R), Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), to appear, arXiv 1711.05080.
  • [FI2] A. Fialowski and K. Iohara, Homology of Lie algebras of orthogonal and symplectic generalized Jacobi matrices, Atti. Acad. Naz. Lincei Rend. Lincei Mat. Appl., to appear, arXiv 1801.00624.
  • [FT] B. L. Feigin and B. L. Tsygan, Cohomologies of Lie Algebras of Generalized Jacobi Matrices, Funct. Anal. Appl. 17 (1983), 86–87.
  • [G] K. R. Goodearl, Completions of regular rings, Math. Ann. 220 (1976), 229–252.
  • [H] I. Halperin, Regular rank rings, Canad. Jour. Math. 17 (1965), 709–719.
  • [HS] P. J. Hilton and U. Stammbach, A Course in Homological Algebra, Grad. Texts in Math. 4, Springer Verlag, 1971.
  • [JM] M. Jimbo and T. Miwa, Solitons and Infinite Dimensional Lie Algebras, Publ. RIMS Kyoto Univ. 19, 1983, 943–1001.
  • [Ka1] M. Karoubi, Matrices de Jacobi, Périodicité de Bott et CC^{\ast}-algèbres, C.R. Acad. Sc. Paris 268, 1969, 109-194.
  • [Ka2] M. Karoubi, Espaces Classifiants en K-Théorie, Trans. AMS 147, 1970, 75-115.
  • [K] Koenigsberger, Carl Gustav Jacob Jacobi, Leipzig, 1904. In it: Lecture Notes by Jacobi, Berlin, 1848.
  • [Lod] J. L. Loday, Cyclic Homology, Grund. math. Wissenschaften 301, Second ed., 1998.
  • [LP] J. L. Loday and C. Procesi, Holmology of Symplectic and Orthogonal Algebras, Adv. in Math. 69 (1988), 93–108.
  • [LQ] J. L. Loday and D. Quillen, Cyclic homology and the Lie algebra homology of matrices, Comment. Math. Helv. 59 (1984), 565–591.
  • [Q] D. Quillen, Rational Homotopy Theory, Ann. Math. 90 (1969), 205–295.
  • [Sa] M. Sato, Soliton Equations as Dynamical Systems on a Infinite Dimensional Grassmann Manifolds, RIMIS Kokyuroku 439 (1981), 30-46.
  • [St] D. Stefan, Hochschild cohomology on Hopf-Galois extensions, Jour. Pure Appl. Alg. 103 (1995), 221–233.
  • [Sz] G. Szegö, Orthogonal Polynomials, Colloq. Publ. 23, Amer. Math. Soc., 1939.
  • [Ts] B. L. Tsygan, Homology of matrix Lie algebras over rings and the Hochschild homology, Russ. Math. Survey 38, 1983, 217–218.
  • [vN] J. von Neumann, Continuous Geometry, Princeton Univ. Press, 1960.